• Tidak ada hasil yang ditemukan

getdoced80. 310KB Jun 04 2011 12:05:09 AM

N/A
N/A
Protected

Academic year: 2017

Membagikan "getdoced80. 310KB Jun 04 2011 12:05:09 AM"

Copied!
33
0
0

Teks penuh

(1)

El e c t ro n ic

Jo ur

n a l o

f P

r o

b a b il i t y

Vol. 15 (2010), Paper no. 63, pages 1938–. Journal URL

http://www.math.washington.edu/~ejpecp/

The center of mass for spatial branching processes and an

application for self-interaction

János Engländer

Abstract

Consider the center of mass of ad-dimensional supercritical branching-Brownian motion. We first show that it is a Brownian motion being slowed down such that it tends to a limiting position almost surely, and that this is also true for a model where branching Brownian motion is modified byattraction/repulsion between particles, where the strength of the attraction/repulsion is given by the parameterγ6=0.

We then put this observation together with the description of the interacting system as viewed from its center of mass, and get our main result in Theorem 16: Ifγ >0 (attraction), then, as

n→ ∞,

2−nZn(dy) w

γ

π

‹d/2

exp€γ|yx|2Šdy, Pxa.s.

for almost allx∈Rd, whereZ(dy)denotes the discrete measure-valued process corresponding

to the interacting branching particle system,⇒w denotes weak convergence, andPxdenotes the law of the particle system conditioned to havexas the limit for the center of mass.

A conjecture is stated regarding the behavior of the local mass in the repulsive case.

We also consider a supercritical super-Brownian motion, and show that, conditioned on survival, its center of mass is a continuous process having an a.s. limit ast→ ∞.

Key words: Branching Brownian motion, super-Brownian motion, center of mass, self-interaction, Curie-Weiss model, McKean-Vlasov limit, branching Ornstein-Uhlenbeck process, spatial branching processes,H-transform.

Department of Mathematics, University of Colorado at Boulder, Boulder, CO-80309, USA.

(2)

AMS 2000 Subject Classification:Primary 60J60; Secondary: 60J80.

(3)

1

Introduction

1.1

Notation

1. Probability:The symbolEwill denote the complement of the eventE, andXY will denote the independent sum of the random variablesX andY.

2. Topology and measures:The boundary of the setBwill be denoted by∂Band the closure of Bwill be denoted by cl(B), that is cl(B):=B∂B; the interior ofBwill be denoted by ˙Band will denote theε-neighborhood ofB. We will also use the notation ˙:={y B: Bε+yB}, whereB+b:={y : ybB}andBt :={x∈Rd : |x|<t}. By abounded rational rectangle

we will mean a setB Rd of the formB =I

I2× · · · ×Id, where Ii is a bounded interval

with rational endpoints for each 1id. The family of all bounded rational rectangles will be denoted byR.

Mf(Rd) andM1(Rd) will denote the space of finite measures and the space of probability

measures, respectively, onRd. For µ ∈ M

f(Rd), we definekµk := µ(Rd). |B| will denote

the Lebesgue measure ofB. The symbols “w” and“v ” will denote convergence in the weak topology and in the vague topology, respectively.

3. Functions:For f,g>0, the notation f(x) =O(g(x))will mean that f(x)C g(x)if x> x0 with somex0∈R,C >0; fgwill mean that f/gtends to 1 given that the argument tends to an appropriate limit. For N R functions the notation f(n) = Θ(g(n)) will mean that c f(n)/g(n)C n, with some c,C >0.

4. Matrices:The symbolId will denote the d-dimensional unit matrix, and r(A)will denote the rank of a matrixA.

5. Labeling:In this paper we will often talk about the ‘ithparticle’ of a branching particle system. By this we will mean that we label the particles randomly, but in a way that does not depend on their spatial position.

1.2

A model with self-interaction

Consider a dyadic (i.e. precisely two offspring replaces the parent) branching Brownian motion (BBM) inRd with unit time branching and with the following interaction between particles: if Z

denotes the process andZti is theithparticle, thenZti ‘feels’ the drift

1 nt

X

1≤jnt

γ·Ztj− ·,

whereγ6=0 , that is the particle’s infinitesimal generator is

1 2∆ +

1

nt

X

1≤jnt

γ·

Ztjx

· ∇. (1.1)

(Here and in the sequel,nt is a shorthand for 2⌊t⌋, where⌊t⌋is the integer part oft.) Ifγ >0, then

(4)

To be a bit more precise, we can define the process by induction as follows. Z0 is a single particle at the origin. In the time interval[m,m+1)we define a system of 2m interacting diffusions, starting at the position of their parents at the end of the previous step (at timem−0) by the following system of SDE’s:

dZti=dWtm,i+ γ 2m

X

1≤j≤2m

(ZtjZti)dt; i=1, 2, . . . , 2m, (1.2)

whereWm,i,i=1, 2, . . . , 2m;m=0, 1, ... are independent Brownian motions.

Remark 1 (Attractive interaction). If there were no branching and the interval [m,m+1) were extended to [0,), then for γ > 0 the interaction (1.2) would describe the ferromagnetic Curie-Weiss model, a model appearing in the microscopic statistical description of a spatially homogeneous gas in a granular medium. It is known that asm→ ∞, a Law of Large Numbers, theMcKean-Vlasov limitholds and the normalized empirical measure

ρm(t):=2−m

2m X

i=1

δZi t

tends to a probability measure-valued solution of

∂tρ=

1 2∆ρ+

γ

2∇ · ρf

ρ,

where (x):=RRd|xy|

2ρ(d y). (See p. 24 in[8]and the references therein.)

Remark 2(More general interaction). It seems natural to replace the linearity of the interaction by a more general rule. That is, to define and analyze the system where (1.2) is replaced by

dZti=dWtm,i+2−m X 1≤j≤2m

g(|ZtjZti|) Z

j tZti

|ZtjZi t|

dt; i=1, 2, . . . , 2m,

where the function g :R+ R has some nice properties. (In this paper we treat the g(x) =γx

case.) This is part of a future project with J. Feng. ⋄

1.3

Existence and uniqueness

Notice that the 2minteracting diffusions on[m,m+1)can be considered as a single 2md-dimensional Brownian motion with linear (and therefore Lipschitz) driftb:R2md

→R2md

:

b x1,x2, ...,xd,x1+d,x2+d, ...,x2d, ...,x1+(2m1)d,x2+(2m1)d, ...,x2md) =:γ(β1,β2, ...,β2md)T,

where

βk=2−m(xbk+xbk+d+...+xbk+(2m1)d)−xk, 1≤k≤2md,

(5)

1.4

Results on the self-interacting model

We are interested in the long time behavior of Z, and also whether we can say something about the number of particles in a given compact set for nlarge. In the sequel we will use the standard notationg,Zt=Zt,g:=Pnt

i=1g(Z

i t).

In this paper we will first show that Z asymptotically becomes a branching Ornstein-Uhlenbeck process (inward for attraction and outward for repulsion), but

1. the origin is shifted to a random point which hasd-dimensional normal distributionN(0, 2Id), and

2. the Ornstein-Uhlenbeck particles are not independent but constitute a system with a degree of freedom which is less than their number by precisely one.

The main result of this articleconcerns the local behavior of the system: we will prove a scaling limit theorem (Theorem 16) for the local mass whenγ >0 (attraction), and formulate and motivate a conjecture (Conjecture 18) whenγ <0 (repulsion).

1.5

The center of mass for supercritical super-Brownian motion

In Lemma 6 we will show that Zt := n1 t

Pnt

i=1Zti, the center of mass for Z satisfies limt→∞Zt =N,

whereN ∼ N(0, 2Id). In fact, the proof will reveal that Zmoves like a Brownian motion, which is nevertheless slowed down tending to a final limiting location (see Lemma 6 and its proof).

Since this is also true forγ=0 (BBM with unit time branching and no self-interaction), our first nat-ural question is whether we can prove a similar result for the supercritical super-Brownian motion.

Let X be the (1

2∆,β,α;R

d)-superdiffusion with α,β > 0 (supercritical super-Brownian motion).

Hereβ is the ‘mass creation parameter’ or ‘mass drift’, whileα > 0 is the ‘variance (or intensity) parameter’ of the branching — see[6]for more elaboration and for a more general setting.

Let Pµ denote the corresponding probability when the initial finite measure isµ. (We will use the abbreviationP:=0.) Let us restrictΩto the survival set

S:={ω∈Ω|Xt(ω)>0, t>0}.

Sinceβ >0,Pµ(S)>0 for allµ6=0.

It turns out that on the survival set the center of mass forX stabilizes:

Theorem 3. Let α,β > 0 and let X denote the center of mass process for the (12∆,β,α;Rd)

-superdiffusion X , that is let

X:= 〈id,X〉 kXk ,

wheref,X:=RRd f(x)X(dx)andid(x) = x. Then, on S, X is continuous and converges P-almost

(6)

Remark 4. A heuristic argument for the convergence is as follows. Obviously, the center of mass is invariant underH-transforms1 wheneverH is spatially (but not temporarily) constant. LetH(t):= eβt. ThenXH is a(1

2∆, 0,e

βtα;Rd)-superdiffusion, that is, a critical super-Brownian motion with

a clock that is slowing down. Therefore, heuristically it seems plausible thatXH, the center of mass

for the transformed process stabilizes, because after some large timeT, if the process is still alive, it behaves more or less like the heat flow (eβtα is small), under which the center of mass does not

move.

1.6

An interacting superprocess model

The next goal is to construct and investigate the properties of a measure-valued process with repre-sentative particles that are attracted to or repulsed from its center of mass.

There is one work in this direction we are aware of: motivated by the present paper, H. Gill [9] has constructed very recently asuperprocess with attraction to its center of mass. More precisely, Gill constructs a supercritical interacting measure-valued process with representative particles that are attracted to or repulsed from its center of mass using Perkins’shistorical stochastic calculus.

In the attractive case, Gill proves the equivalent of our Theorem 16 (see later): on S, the mass normalized process converges almost surely to the stationary distribution of the Ornstein-Uhlenbeck process centered at the limiting value of its center of mass; in the repulsive case, he obtains sub-stantial results concerning the equivalent of our Conjecture 18 (see later), using[7]. In addition, a version of Tribe’s result[12]is presented in[9].

2

The mass center stabilizes

Returning to the discrete interacting branching system, notice that

1 nt

X

1≤jnt

ZtjZti

=ZtZti, (2.1)

and sothe net attraction pulls the particle towards the center of mass(net repulsion pushes it away from the center of mass).

Remark 5. The reader might be interested in the very recent work[11], where a one dimensional particle system is considered with interaction via the center of mass. There is a kind of attraction towards the center of mass in the following sense: each particle jumps to the right according to some common distributionF, but the rate at which the jump occurs is a monotone decreasing function of the signed distance between the particle and the mass center. Particles being far ahead slow down,

while the laggards catch up.

Since the interaction is in fact through the center of mass, the following lemma is relevant:

Lemma 6(Mass center stabilizes). There is a random variable N∼ N(0, 2Id)such thatlimt→∞Zt=

N a.s.

(7)

Proof. LetmN. Fort[m,m+1)there are 2m particles moving around and particleZi

t’s motion

is governed by

dZti=dWtm,i+γ(ZtZti)dt.

Since 2mZt=P2i=m1Zti, we obtain that

dZt=2−m

2m X

i=1

dZti=2−m 2m X

i=1

dWtm,i+ γ 2m 2

mZ t

2m X

i=1 Zti

!

dt =2−m 2m X

i=1

dWtm,i.

So, for 0τ <1,

Zm+τ=Zm+2−m

2m

M

i=1

Wτm,i =:Zm+2−m/2B(m)(τ), (2.2)

where we note thatB(m)is a Brownian motion on[m,m+1). Using induction, we obtain that2

Zt=B(0)(1)⊕

1 p

2B (1)(1)

⊕ · · · ⊕ 1 2k/2B

(k)(1)

· · · ⊕p 1

2⌊t⌋−1

B(⌊t⌋−1)(1)p1 ntB

(⌊t⌋)(τ), (2.3)

whereτ:=t− ⌊t⌋.

By Brownian scalingW(m)(·):=2−m/2B(m)(2m·), m≥1 are (independent) Brownian motions. We have

Zt=W(0)(1)⊕W(1)

1

2

⊕ · · · ⊕W(⌊t⌋−1)

1

2⌊t⌋−1

W(⌊t⌋)

τ

nt

.

By the Markov property, in fact

Zt=Wc

1+1

2+· · ·+ 1

2⌊t⌋−1 +

τ

nt

,

whereWcis a Brownian motion (the concatenation of theW(i)’s), and sinceWchas a.s. continuous paths, limt→∞Zt=Wc(2), a.s.

For another proof see the remark after Lemma 9.

Remark 7. It is interesting to note thatZ is in fact a Markov process. To see this, let{Ft}t0 be the canonical filtration for Z and{Gt}t≥0 the canonical filtration for Z. SinceGs ⊂ Fs, it is enough to

check the Markov property withGs replaced byFs.

Assume first 0s< t,s=t=:m. Then the distribution of Zt conditional onFsis the same as conditional on Zs, because Z itself is a Markov process. But the distribution ofZt only depends on Zs throughZs, as

Zt d

=ZsW(2

m)ts

2m

, (2.4)

whatever Zs is. That is, P(Zt∈ · | Fs) =P(Zt ∈ · |Zs) = P(Zt ∈ · |Zs). Note that this is even true whensNandt=s+1, becauseZt=Zt0.

Assume now thats<t=:m. Then the equation P(Zt ∈ · | Fs) = P(Zt ∈ · | Zs), is obtained by

conditioning successively onm,m−1, ...,⌊s⌋+1,s. ⋄

(8)

We will also need the following fact later.

Lemma 8. The coordinate processes of Z are independent one dimensional interactive branching pro-cesses of the same type as Z.

We leave the simple proof to the reader.

3

Normality via decomposition

Let againm:=t⌋. We will need the following result.

Lemma 9 (Decomposition). In the time interval [m,m + 1) the d · 2m-dimensional process (Zt1,Zt2, ...,Zt2m) can be decomposed into two components: a d-dimensional Brownian motion and an independent d(2m1)-dimensional Ornstein-Uhlenbeck process with parameterγ.

More precisely, each coordinate process (as a 2m-dimensional process) can be decomposed into two components: a one-dimensional Brownian motion in the direction(1, 1, ..., 1)and an independent(2m 1)-dimensional Ornstein-Uhlenbeck process with parameter γ in the ortho-complement of the vector (1, 1, ..., 1).

Furthermore,(Zt1,Zt2, ...,Znt

t )is d nt-dimensional joint normal for all t≥0.

Proof of lemma.By Lemma 8, we may assume thatd=1. We prove the statement by induction.

(i) Form=0 it is trivial.

(ii) Suppose that the statement is true form−1. Consider the timemposition of the 2m−1particles (Zm1,Zm2, ...,Zm2m−1)‘just before’ the fission. At the instant of the fission we obtain the 2m-dimensional vector

(Zm1,Zm1,Zm2,Zm2, ...,Zm2m−1,Zm2m−1),

which has the same distribution on the 2m−1 dimensional subspace S:={xR2m

|x1=x2,x3= x4, ...,x2m1=x2m}

of R2m as the vector p2(Z1

m,Zm2, ...,Z2 m−1

m ) on R2

m−1

. Since, by the induction hypothesis, (Zm1,Zm2, ...,Zm2m−1) is normal, the vector formed by the particle positions ‘right after’ the fission is a 2m-dimensional degenerate normal. (The reader can easily visualize this form=1: the distribu-tion of(Z11,Z11)is clearlyp2 times the distribution of a Brownian particle at time 1, i.e. N(0,p2) on the line x1=x2.)

Since the convolution of normals is normal, therefore, by the Markov property, it is enough to prove the statement when the 2m particles start at the origin and the clock is reset: t[0, 1).

Define the 2m-dimensional processZ∗on the time intervalt[0, 1)by Zt∗:= (Z1t,Zt2, ...,Z2tm),

starting at the origin. Because the interaction between the particles attracts the particles towards the center of mass,Z∗is a Brownian motion with drift

γ”(Zt,Zt, ...,Zt)−(Zt1,Z

2

t, ...,Z

2m

t )

—

(9)

Notice that this drift is orthogonal to the vector3 v:= (1, 1, ..., 1), that is, the vector (Zt,Zt, ...,Zt) is nothing but the orthogonal projection of(Zt1,Zt2, ...,Zt2m)to the line ofv. This observation imme-diately leads to the following decomposition. The process Z∗can be decomposed into two compo-nents:

• the component in the direction ofvis a Brownian motion

• in S, the ortho-complement of v, it is an independent Ornstein-Uhlenbeck process with pa-rameterγ.

Remark 10. (i) The proof also shows that (Zt1,Z2t, ...,Znt

t ) given Zs is a.s. joint normal for all

t>s≥0.

(ii) Consider the Brownian component in the decomposition appearing in the proof. Since, on the other hand, this coordinate is 2m/2Zt, using Brownian scaling, one obtains a slightly different way of

seeing thatZt stabilizes at a position which is distributed as the time 1+2−1+2−2+...+2−m+...=2 value of a Brownian motion. (The decomposition shows this ford =1 and then it is immediately

upgraded to generald by independence.) ⋄

Corollary 11 (Asymptotics for finite subsystem). Let k ≥ 1 and consider the subsystem (Zt1,Zt2, ...,Ztk), t m0 for m0 := log2k+1. (This means that at time m0 we pick k particles and at every fission replace the parent particle by randomly picking one of its two descendants.) Let the real numbers c1, ...,cksatisfy

k

X

i=1 ci=0,

k

X

i=1

ci2=1. (3.1)

DefineΨt = Ψ(c1,...,ck)

t :=

Pk

i=1ciZti and note thatΨt is invariant under the translations of the

coordi-nate system. LetLt denote its law.

For every k≥1and c1, ...,ck satisfying (3.1),Ψ(c1,...,ck)is the same d-dimensional Ornstein-Uhlenbeck

process corresponding to the operator1/2∆γ∇ ·x, and in particular, forγ >0,

lim

t→∞Lt=N

0, 1 2γId

.

For example, takingc1=1/p2,c2=1/p2, we obtain that when viewed from a tagged particle’s position, any given other particle moves asp2 times the above Ornstein-Uhlenbeck process. Proof. By independence (Lemma 8) it is enough to consider d = 1. For m fixed, consider the decomposition appearing in the proof of Lemma 9 and recall the notation. By (3.1), whatever mm0is, the 2m dimensional unit vector

(c1,c2, ...,ck, 0, 0, ..., 0)

is orthogonal to the 2m dimensional vectorv. This means thatΨ(c1,...,ck) is a one dimensional

pro-jection of the Ornstein-Uhlenbeck component ofZ∗, and thus it is itself a one dimensional Ornstein-Uhlenbeck process (with parameterγ) on the unit time interval.

Now, although as m grows, the Ornstein-Uhlenbeck components of Z∗ are defined on larger and larger spaces (S ⊂ R2m

is a 2m−1 dimensional linear subspace), the projection onto the direction of(c1,c2, ...,ck, 0, 0, ..., 0) is always the same one dimensional Ornstein-Uhlenbeck process, i.e. the

different unit time ‘pieces’ ofΨ(c1,...,ck)obtained by those projections may be concatenated.

(10)

4

The interacting system as viewed from the center of mass

Recall that by (2.2) the interaction has no effect on the motion of Z. Let us see now how the interacting system looks like when viewed fromZ.

4.1

The description of a single particle

Using our usual notation, assume thatt[m,m+1)and letτ:=t− ⌊t. When viewed from Z, the relocation4 of a particle is governed by

d(Z1tZt) =dZt1−dZt=dWtm,1−2−m

2m X

i=1

dWtm,iγ(Z1tZt)dt.

So ifY1:=Z1Z, then

dYt1=dWtm,1−2−m 2m

X

i=1

dWtm,iγYt1dt.

Clearly,

Wτm,12−m 2m M

i=1

Wτm,i = 2m M

i=2

2−mWτm,i(12−m)Wτm,1;

and, by a trivial computation, the right hand side is a Brownian motion with mean zero and variance (12−m)τId :=σ2mτId. That is,

dYt1=σmdWf m,1

tγYt1dt,

whereWfm,1is a standard Brownian motion.

We have thus obtained that on the time interval [m,m+1), Y1 corresponds to the Ornstein-Uhlenbeck operator

1 2σ

2

m∆−γx· ∇. (4.1)

Since formlargeσmis close to one, the relocation viewed from the center of mass isasymptotically

governed by an O-U process corresponding to 1

2∆−γx· ∇.

Remark 12(Asymptotically vanishing correlation between driving BM’s). LetfWm,i,k be thekth co-ordinate of theithBrownian motion:Wfm,i= (Wfm,i,k,k=1, 2, ...,d)andBm,i,kbe thekthcoordinate ofWm,i. For 1i6= j2m, we have

E”σmWfτm,i,k·σmWfτm,j,k

— =

E  

Bτm,i,k2−m 2m M

r=1 Bτm,r,k

 

Bτm,j,k2−m 2m M

r=1 Bmτ,r,k

 

 =

(11)

that is, fori6= j,

E”fWτm,i,kWfτm,j,—= δkℓ

2m−1τ. (4.2)

Hencethe pairwise correlation tends to zeroas t→ ∞(recall thatm=tandτ=tm[0, 1)).

And of course, for the variances we have

E”Wfτm,i,kWfτm,i,—=δkℓ·τ, for 1≤i≤2m. ⋄ (4.3)

4.2

The description of the system; the ‘degree of freedom’

Fixm1 and fort [m,m+1)let Yt:= (Yt1, ...,Yt2m)T, where()T denotes transposed. (This is a vector of length 2m where each component itself is ad dimensional vector; one can actually view it as a 2m×dmatrix too.) We then have

dYt=σmdWf

(m)

tγYtdt,

where

f

W(m)=€Wfm,1, ...,Wfm,2mŠT

and

f

Wτm,i=σm1

 

Wτm,i2−m 2m M

j=1 Wτm,j

 

, i=1, 2, ..., 2m

are mean zero Brownian motions with correlation structure given by (4.2)-(4.3).

Just like at the end of Subsection 1.2, we can considerY as a single 2md-dimensional diffusion. Each of its components is an Ornstein-Uhlenbeck process with asymptotically unit diffusion coefficient.

By independence, it is enough to consider thed=1 case, and so from now on, in this subsection we assume thatd=1.

Let us first describe the distribution ofWft(m)fort0 fixed. Recall that{Wsm,i, s0; i=1, 2, ..., 2m} are independent Brownian motions. By definition,Wft(m)is a 2m-dimensional multivariate normal:

f

Wt(m)=σm

        

1−2−m −2−m ... −2−m −2−m 12−m ... 2−m

. . .

−2−m 2−m ... 12−m

        

Wt(m)=:σm−1A(m)Wt(m),

whereWt(m)= (Wtm,1, ...,Wtm,2m)T, yielding

dYt=A(m)dWt(m)γYtdt.

Since we are viewing the system from the center of mass, Wft(m) is asingular multivariate normal and thusY is a degenerate diffusion. The ‘true’ dimension ofWft(m)is r(A(m)).

(12)

Proof. We will simply writeAinstead ofA(m). Since the columns ofAadd up to zero, the matrixA is not of full rank: r(A)2m1. On the other hand,

2mA+         

1 1 ... 1 1 1 ... 1

. . .

1 1 ... 1

        

=2mI,

whereIis the 2m-dimensional unit matrix, and so by subadditivity,

r(A) +1=r(2mA) +12m.

By Lemma 13,Wft(m)is concentrated onS, and there the vectorWft(m) hasnon-singularmultivariate normal distribution.5 What this means is that even thoughWfm,1, ...,Wfm,2m are not independent, their ‘degree of freedom’ is 2m−1, i. e. the 2m-dimensional vector Wft(m) isdetermined by2m−1 independent components(corresponding to 2m1 principal axes).

5

Asymptotic behavior

5.1

Conditioning

How can one put together thatZttends to a random final position with the description of the system ‘as viewed from Zt?’ The following lemma is the first step in this direction.

Lemma 14(Independence). LetT be the tailσ-algebra of Z.

1. For t≥0, the random vector Yt is independent of the path{Zs}st.

2. The process Y = (Yt;t≥0)is independent ofT.

Proof. In both parts we will refer to the following fact. Letst,s[mb,mb+1);t[m,m+1)with

b

mm. Since the random variables Z1t,Zt2, ...,Zt2m are exchangeable, thus, denoting bn:=2mb,n:= 2m, the vectors Zt andZs1Zs are uncorrelated for 0st. Indeed, by Lemma 8 we may assume thatd=1 and then

E[Zt·€Zs1ZsŠ] =E  Z

1

t +Z

2

t +...+Z n t

n · Z

1

s

Zs1+Zs2+...+Zsbn b

n

! =

1 nE

€

Zt1·Zs1Š+n−1 n E

€

Zs1·Zt bn nbnE

€

Zt1·Zsbn(n−1) nbn E

€

Zt2·Zs1Š=0.

(Of course the index 1 can be replaced byifor any 1i≤2m.)

(13)

Part (1): First, for anyt>0, the (d·2m-dimensional) vectorYtis independent of the (d-dimensional) vectorZt, because thed(2m+1)-dimensional vector

(Zt,Z1tZt,Z2tZt, . . . ,Z2 m

tZt)T

is normal (since it is a linear transformation of the d · 2m dimensional vector (Zt1,Zt2, . . . ,Zt2m)T, which is normal by Lemma 9), and so it is sufficient to recall that Zt and

ZtiZtare uncorrelated for 1≤i≤2m.

To complete the proof of (a), recall (2.2) and (2.3) and notice that the conditional distribution of {Zs}st givenFtonly depends on its starting pointZt, as it is that of a Brownian path appropriately slowed down, whateverYt(or, equivalently, whatever Zt=Yt+Zt) is. Since, as we have seen,Yt is independent ofZt, we are done.

Part (2): LetA∈ T. By Dynkin’s Lemma, it is enough to show that(Yt1, ...,Ytk) is independent of

Afor 0 t1 < ... < tk and k 1. Since A∈ T ⊂σ(Zs;s tk+1), it is sufficient to show that (Yt

1, ...,Ytk)is independent of{Zs}stk+1.

To see this, similarly as in Part (1), notice that the conditional distribution of{Zs}stk+1givenFtk+1

only depends on its starting pointZtk+1, as it is that of a Brownian path appropriately slowed down,

whatever the vector (Yt

1, ...,Ytk) is. If we show that (Yt1, ...,Ytk) is independent of Ztk+1, we are

done.

To see why this latter one is true, one just have to repeat the argument in (a), using again normality6 and recalling that the vectorsZt andZsiZs are uncorrelated. ƒ

Remark 15 (Conditioning on the final position of Z). Let N := limt→∞Zt (recall that N

N(0, 2Id)) and

Px(·):=P(· |N= x).

By Lemma 14, Px(Yt ∈ ·) =P(Yt ∈ ·)for almost all x Rd. It then follows that the decomposition

Zt = ZtYt as well as the result obtained for the distribution ofY in subsections 4.1 and 4.2 are

true underPx too, for almost allx ∈Rd.

5.2

Main result and a conjecture

So far we have obtained that on the time interval [m,m+1), Y1 corresponds to the Ornstein-Uhlenbeck operator

1 2σ

2

m∆−γx· ∇,

whereσm→1 asm→ ∞, with asymptotically vanishing correlation between the driving Brownian

motions; that

dYt=A(m)dW

(m)

tγYtdt,

where{Wsm,i, s≥0; i=1, 2, ..., 2m}are independent Brownian motions and r(A(m)) =2m−1, and finally, the independence of the center of mass and the relative motions as in Lemma 14.

6We now need normality for finite dimensional distributions and not just for one dimensional marginals, but this is

(14)

We now go beyond these preliminary results and state a theorem (the main result of this article) and a conjecture on the local behavior of the system. Recall that one can considerZnas an element

ofMf(Rd)by putting unit point mass at the site of each particle; with a slight abuse of notation we

will writeZn(dy). Let{Px,x Rd

}be as in Remark 15. Our main result is as follows.

Theorem 16(Scaling limit for the attractive case). Ifγ >0, then, as n→ ∞,

2−nZn(dy) w

γ

π

‹d/2

exp€γ|yx|2Šdy, Px−a.s. (5.1)

for almost all xRd. Consequently,

2−nE Zn(dy)w (y)dy, (5.2) where

(·) =€π(4+γ−1)Š−d/2exp

–

−| · |2 4+γ−1

™

.

Remark 17.

(i) The proof of Theorem 16 will reveal that actually

2−tnZ tn(dy)

w

γ

π

‹d/2

exp€γ|yx|2Šdy, Px a.s.

holds for any given sequence{tn}with tn↑ ∞asn→ ∞. This, of course, is still weaker than P-a.s. convergence as t → ∞, but one can probably argue, using the method of Asmussen and Hering, as in Subsection 4.3 of[4]to upgrade it to continuous time convergence. Nev-ertheless, since our model is defined with unit time branching anyway, we felt satisfied with (5.1).

(ii) Notice that , which is the limiting density of the intensity measure, is the density for N

0,

2+21γ

Id

. This is the convolution ofN 0, 2Id, representing the randomness of

the final position of the center of mass (c.f. Lemma 6) and N 0,1 2γ

Id, representing the final distribution of the mass scaled Ornstein-Uhlenbeck branching particle system around its center of mass (c.f. (5.1)). For strong attraction, the contribution of the second term is

negligible.

Conjecture 18(Dichotomy for the repulsive case). Letγ <0.

1. If|γ| ≥ log 2d , then Z suffers local extinction:

Zn(dy)v 0, Pa.s.

2. If|γ|< log 2d , then

(15)

5.3

Discussion of the conjecture

The intuition behind the phase transition at log 2/d is as follows. For a branching diffusion onRd

with motion generator L, smooth nonzero spatially dependent exponential branching rate β(·) 0 and dyadic branching, it is known (see Theorem 3 in [5]) that either local extinction or local exponential growth takes place according to whetherλc≤0 orλc>0, whereλc=λc(L+β)is the

generalized principle eigenvalue7of L+β onRd. In particular, forβ

B>0, the criterion for local exponential growth becomesB>|λc(L)|, whereλc(L)≤0 is the generalized principle eigenvalue of

L, which is also the ‘exponential rate of escape from compacts’ for the diffusion corresponding toL. The interpretation of the criterion in this case is that a large enough mass creation can compensate the fact that individual particles drift away from a given bounded set. (Note that if Lcorrespond to a recurrent diffusion thenλc(L) =0. )

In our case, the situation is similar, withλc=for the outward Ornstein-Uhlenbeck process, taking

into account that for unit time branching, the role of B is played by log 2. The condition for local exponential growth should therefore be log 2>d|γ|.

The scaling 2−ned|γ|n comes from a similar consideration, noting that in our unit time branching setting, 2n replaces the termeβt appearing in the exponential branching case, whileeλc(L)tbecomes

eλc(L)n=edγn.

For a continuous time result analogous to (2) see Example 11 in[4]. Note that since the rescaled (vague) limit of Zn(dy) is translation invariant (i.e. Lebesgue), the final position of the center of mass plays no role.

Although we will not prove Conjecture 18, we will discuss some of the technicalities in section 8.

6

Proof of Theorem 3

Sinceα,βare constant, the branching is independent of the motion, and thereforeN defined by

Nt:=eβtkXtk

is a nonnegative martingale (positive onS) tending to a limit almost surely. It is straightforward to check that it is uniformly bounded in L2 and is therefore uniformly integrable (UI). Write

Xt=

eβtid,Xt eβtkX

tk

= e

βt

〈id,Xt Nt .

We now claim that N > 0 a.s. on S. Let A:= {N = 0}. Clearly S A, and so if we show that P(A) = P(S), then we are done. As is well known, P(S) = eβ/α. On the other hand, a

standard martingale argument (see the argument after formula (20) in[3]) shows that 0u(x):= −logPδx(A)must solve the equation

1

2∆u+βuαu 2=0,

but sincePδx(A) =P(A)constant, therefore−logPδx(A)solvesβuαu

2=0. SinceNis UI, no mass

is lost in the limit, givingP(A)<1. Sou>0, which in turn implies thatlogPδ

(16)

Once we know thatN>0 a.s. onS , it is enough to focus on the term eβtid,Xt: we are going to show that it converges almost surely. Clearly, it is enough to prove this coordinate-wise.

Recall a particular case of theH-transform for the(L,β,α;Rd)-superdiffusionX (see Appendix B in

[7]):

Lemma 19. Define XH by

XHt :=H(·,t)Xt ‚

that is, dX

H t

dXt

=H(·,t) Œ

, t 0. (6.1)

If X is an(L,β,α;Rd)-superdiffusion, and H(x,t):=eλth(x),where h is a positive solution of (L+

β)h=λh, then XH is a(L+ahh· ∇, 0,eλtαh;Rd)-superdiffusion.

In our case β(·) β. So choosing h(·) 1 and λ = β, we have H(t) = eβt and XH is a (1

2∆, 0,e

βtα;Rd)-superdiffusion, that is, a critical super-Brownian motion with a clock that is

slow-ing down. Since, as noted above, it is enough to prove the convergence coordinate-wise, we can assume thatd=1. One can write

eβtid,Xt=id,XHt .

Let{Ss}s0 be the ‘expectation semigroup’ forX, that is, the semigroup corresponding to the oper-ator 12∆ +β. The expectation semigroup{SsH}s≥0 forXH satisfies Ts :=SsH =e

βs

Ss and thus it

corresponds to Brownian motion. In particular then

Ts[id] =id. (6.2)

(One can pass from bounded continuous functions to f := id by defining f1 := f1x>0 and f2 := f1x0, then noting that by monotone convergence, Eδxfi,X

H

t 〉= Exfi(Wt) ∈(−∞,∞), i = 1, 2,

whereW is a Brownian motion with expectationE, and finally taking the sum of the two equations.) ThereforeM:=id,XHis a martingale:

Eδ

x Mt| Fs

=Eδ

x €

〈id,XtH〉 | FsŠ=

EXsid,XHt = Z

R

Eδy〈id,X H tX

H s (dy) =

Z

R

y XsH(dy) =Ms.

We now show that M is UI and even uniformly bounded in L2, verifying its a.s. convergence, and that of the center of mass. To achieve this, definegn by gn(x) =|x| ·1{|x|<n}. Then we have

Eid,XtH2=E|〈id,XtH〉|2E〈|id|,XtH2,

and by the monotone convergence theorem we can continue with

= lim

n→∞Egn,X

H t

2.

Sincegn is compactly supported, there is no problem to use the moment formula and continue with

= lim

n→∞

Z t

0

ds eβsδ0,Ts[αg2n]〉=αnlim→∞

Z t

0

(17)

Recall that{Ts;s0}is the Brownian semigroup, that is,Ts[f](x) =Exf(Ws), whereW is Brown-ian motion. Sincegn(x)≤ |x|, therefore we can trivially upper estimate the last expression by

α

Z t

0

ds eβsE0(W2

s ) =α

Z t

0

ds seβs=α

‚

1eβt

β2 − t eβt

β

Œ

< α β2.

Since this upper estimate is independent oft, we are done:

sup

t≥0

Eid,XtH2 α

β2.

Finally, we show that X has continuous paths. To this end we first note that we can (and will) consider a version of X where all the paths are continuous in the weak topology of measures. We now need a simple lemma.

Lemma 20. Let{µt, t ≥0}be a family inMf(Rd)and assume that t0>0andµt w

µt0 as tt0. Assume furthermore that

C=Ct0,ε:=cl   

t[0+ε

t=t0−ε

supp(µt)

  

is compact with someε >0. Let f :Rd

→Rd be a continuous function. Thenlim

tt0〈f,µt〉=〈f,µt0〉.

Proof of Lemma 20. First, if f = (f1, ...,fd) then all fi are Rd → R continuous functions and

limtt0f,µt〉 = 〈f,µt0〉 simply means that limtt0〈fi,µt〉 = 〈fi,µt0〉. Therefore, it is enough to prove the lemma for anRd

→R continuous function. Let k be so large that C Ik := [k,k]d. Using a mollified version of1[−k,k], it is trivial to construct a continuous function bf =:Rd→Rsuch that bf = f on Ik and bf =0 onRd

\I2k. Then,

lim

tt0〈

f,µt〉=tlim

t0〈 b

f,µt〉=〈fb,µt0〉=〈f,µt0〉,

since bf is a bounded continuous function. ƒ

Returning to the proof of the theorem, let us invoke the fact that for

Cs(ω):=cl [

zs

supp(Xz(ω)) !

,

we have P(Cs is compact) = 1 for all fixed s 0 (compact support property; see [6]). By the monotonicity ins, there exists anΩ1ΩwithP(Ω1) =1 such that forω∈Ω1,

Cs(ω)is compacts0.

Letω∈Ω1and recall that we are working with a continuous path version ofX. Then letting f :=id and µt = Xt(ω), Lemma 20 implies that for t0 > 0, limtt0〈id,Xt(ω)〉 = 〈id,Xt0(ω)〉. The right

(18)

7

Proof of Theorem 16

Before we prove (5.1), we note the following. Fixx Rd. Abbreviate

ν(x)(dy):= γ

π

‹d/2

exp€γ|yx|2Šdy.

Since 2−nZn(dx),ν(x)∈ M1(Rd), therefore, as is well known8, 2nZ n(dx)

w

ν(x)is in fact equiva-lent to

g∈ E : 2−ng,Zn〉 → 〈g,ν(x)〉,

whereE is any given family of bounded measurable functions withν(x)-zero (Lebesgue-zero) sets of discontinuity that is separating forM1(Rd).

In fact, one can pick acountableE, which, furthermore, consists of compactly supported functions. Such anE is given by the indicators of sets inR.

Fix such a familyE. SinceE is countable, in order to show (5.1), it is sufficient to prove that for almost allx ∈Rd,

Px(2−ng,Zn〉 → 〈g,ν(x)〉) =1, g∈ E. (7.1) We will carry out the proof of (7.1) in several subsections.

7.1

Putting

Y

and

Z

together

The following remark is for the interested reader familiar with [4] only. It can be skipped without any problem.

Remark 21. Once we have the description ofY as in Subsections 4.1 and 4.2 and Remark 15, we can try to put them together with the Strong Law of Large Numbers for the local mass from[4]for the processY.

If the components of Y were independent and the branching rate were exponential, Theorem 6 of[4]would be readily applicable. However, since the 2m components of Y are not independent (as we have seen, their degree of freedom is 2m1) and since, unlike in [4], we now have unit time branching, the method of [4] must be adapted to our setting. The reader will see that this

adaptation requires quite a bit of extra work.

Let fe(·) = feγ(·):= (πγ)d/2exp€γ| · |2Š, and note that ef is the density forN(0,(2γ)−1Id). We now claim that in order to show (7.1), it is enough to prove that for almost allx,

Px(2−ng,Yn〉 → 〈g,ef) =1, g∈ E. (7.2) This is because

lim

n→∞2 −n

g,Zn〉= lim

n→∞2 −n

g,Yn+Zn〉= lim

n→∞2 −n

g(·+Zn),Yn〉=I+I I,

where

I:= lim

n→∞2 −n

g(·+x),Yn

(19)

and

I I:= lim

n→∞2 −n

g(·+Zn)−g(·+x),Yn〉.

Now, (7.2) implies that for almost all x, I =g(·+x),ef(·) Px-a.s., while the compact support of g, and Heine’s Theorem yields thatI I=0, Pxa.s. Hence, limn→∞2−ng,Zn〉=〈g(·+x),ef(·)〉=

g(·),fe(· −x), Px-a.s., giving (7.1).

Next, let us see how (5.1) implies (5.2). Let g be continuous and bounded. Since 2−nZn,g〉 ≤ kgk, it follows by bounded convergence that

lim

n→∞E2 −n

Zn,g〉=

Z

Rd

Ex 

lim

n→∞2 −n

Zn,g

‹

Q(dx) = Z

Rd

g(·),fe(· −x)Q(dx),

whereQ∼ N(0, 2Id). Now, if fb∼ N(0, 2Id)then, since ∼ N

0,

2+21γ

Id

, it follows that

= bf ∗ef and Z

Rd

g(·),fe(· −x)Q(dx) =g(·),,

yielding (5.2).

Next, notice that it is in fact sufficient to prove (7.2)under P instead of Px. Indeed, by Lemma 14,

Px 

lim

n→∞2 −n

g,Yn〉=〈g,ef

‹ =P



lim

n→∞2 −n

g,Yn〉=〈g,ef〉 |N=x

‹

=P 

lim

n→∞2 −n

g,Yn〉=〈g,fe〉

‹

.

Let us use the shorthandUn(dy):=2−nYt(dy); in generalUt(dy):= 1

ntYt(dy). With this notation,

our goal is to show that

P(g,Un〉 → 〈g,fe) =1, g∈ E. (7.3)

Now, as mentioned earlier, we may (and will) takeE :=I, whereI is the family of indicators of sets inR. Then, it remains to show that

P ‚

Un(B) Z

B

e f(x)dx

Œ

=1, B∈ R. (7.4)

7.2

Outline of the further steps

Notation 22. In the sequel{Ft}t0will denote the canonical filtration forY, rather than the canon-ical filtration forZ.

The following key lemma (Lemma 23) will play an important role. It will be derived using Lemma 25 and (7.12), where the latter one will be derived with the help of Lemma 25 too. Then, Lemma 23 together with (7.14) will be used to complete the proof of (7.4) and hence, that of Theorem 16.

Lemma 23. Let B⊂Rd be a bounded measurable set. Then,

lim

n→∞

”

Un+mn(B)−E(Un+mn(B)| Fn) —

(20)

7.3

Establishing the crucial estimate (7.12) and the key Lemma 23

LetYni denote the “ith” particle at timen,i=1, 2, ..., 2n. SinceBis a fixed set, in the sequel we will simply writeUninstead ofUn(B). Recall the time inhomogeneity of the underlying diffusion process

and note that by the branching property, we have theclumping decomposition: forn,m1,

Un+m= 2n X

i=1

2−nUm(i), (7.6)

where givenFn, each member in the collection{Um(i): i=1, ..., 2n}is defined similarly to Um but withYmreplaced by the timemconfiguration of the particles starting atYni, i=1, ..., 2n, respectively, and with motion component 1

2σn+k∆−γx·∆in the time interval[k,k+1).

7.3.1 The functionsa andζ

Next, we define two positive functions,aandζon(1,). Here is a rough motivation.

(i) The functiona·will be related (via (7.9) below) to theradial speedof the particle systemY. (ii) The functionζ(·), will be related (via (7.10) below) to thespeed of ergodicityof the underlying

Ornstein-Uhlenbeck process.

Fort>1, define

at :=C0· p

t, (7.7)

ζ(t):=C1logt, (7.8)

whereC0andC1 are positive (non-random) constants to be determined later. Note that

mt:=ζ(at) =C3+C4logt

withC3=C1logC0RandC4=C1/2>0. We will use the shorthand

ℓn:=2⌊mn⌋.

Recall that e is the density for N(0,(2γ)−1Id) and let q(x,y,t) =q(γ)(x,y,t) andq(x, dy,t) = q(γ)(x, dy,t)denote the transition density and the transition kernel, respectively, corresponding to the operator 12γx· ∇. We are going to show below that for sufficiently large C0 and C1, the following holds. For each givenx Rd andB

⊂Rd nonempty bounded measurable set,

P€n0,n0<nN : supp(Yn)Ba

n Š

=1, and (7.9)

lim

t→∞z∈supBt,yB

q(z,y,ζ(t)) e

(y) −1

=0. (7.10)

(21)

Remark 9 in [4]), and thus they can be mimicked in our case for the Ornstein-Uhlenbeck process performed by the particles inY (which corresponds to the operator1

2σm∆−γx·∇on[m,m+1), m≥ 1), despite the fact that the particle motions are now correlated. These expectation calculations lead to the estimate that the growth rate of the support ofY satisfies (7.9) with a sufficiently large C0=C0(γ). The same example shows that (7.10) holds with a sufficiently largeC1=C1(γ).

Remark 24. Denote by ν = νγ ∈ M1(Rd) the distribution of

N(0,(2γ)−1Id). Let B Rd be a

nonempty bounded measurable set. Takingt=anin (7.10) and recalling thatζ(an) =mn,

lim

n→∞zBsupan,yB

q(z,y,mn) e

(y) −1 =0.

Since e is bounded, this implies that for any bounded measurable setBRd,

lim

n→∞zsupB

an

q(z,B,mn)ν(B)=0. (7.11)

We will use (7.11) in Subsection 7.4.

7.3.2 Covariance estimates

Let{Ymi,nj, j =1, ...,ℓn}be the descendants ofYni at timemn+n. SoY

1,j

mn andY

2,k

mn are respectively

the jth and kth particle at timemn+nof the trees emanating from the first and second particles

at time n. It will be useful to control the covariance between 1B(Ym1,nj) and1B(Y

2,k

mn), where B is a

nonempty, bounded open set. To this end, we will need the following lemma, the proof of which is relegated to Section 9 in order to minimize the interruption in the main flow of the argument.

Lemma 25. Let BRd be a bounded measurable set.

(a) There exists a non-random constant K(B)such that if C=C(B,γ):= 3

γ|B|

2K(B), then

P

n large enough andξ,ξe∈Πn, ξ6=ξe:

P(ξmn,ξemnB| Fn)−P(ξmnB| Fn)P(ξemnB| Fn) C n

2n

=1,

whereΠndenotes the collection of thoseℓnparticles, which, start at some time-n location of their

parents and run for (an additional) time mn.

(b) Let C=C(B):=ν(B)(ν(B))2. Then

P –

lim

n→∞ξsupΠ

n Var

1{ξmnB}| Fn

C=0

™ =1.

Remark 26. In the sequel, instead of writingξmn andξemn, we will use the notationY i1,j

mn andY

i2,k

mn

(22)

7.3.3 The crucial estimate (7.12)

LetB⊂Rd be a bounded measurable set andC=C(B,γ)as in Lemma 25. Define

Zi:=

1

ℓn ℓn X

j=1

h

1B(Ymi,j

n)−P(Y i,j

mnB|Y i n)

i

, i=1, 2, ..., 2n.

With the help of Lemma 25, we will establish the following crucial estimate, the proof of which is provided in Section 9.

Claim 27. There exists a non-random constantCb(B,γ)>0such that

P   

X

1≤i6=j≤2n

E”ZiZj| Fn

—

Cb(B,γ)nℓn

2n

X

i=1

E”Zi2| Fn

—

, for large n’s  

=1. (7.12)

The significance of Claim 27 is as follows.

Claim 28. Lemma 25 together with the estimate (7.12) implies Lemma 23.

Proof of Claim 28. Assume that (7.12) holds. By the clumping decomposition under (7.6),

Un+mnE(Un+mn| Fn) =

2n X

i=1 2−n

Um(i)

nE(U (i)

mn| Fn)

.

SinceUm(i)

n=

−1

n

Pℓn j=11B(Y

i,j

mn), therefore

Um(i)

nE(U (i)

mn| Fn) =U (i)

mnE(U (i)

mn|Y i n) =Zi.

Hence,

E”Un+mnE(Un+mn| Fn) —2

| Fn

=E      2n X

i=1 2−n

Um(i)

nE(U (i)

mn| Fn) 

2

| Fn

   =E      2n X

Referensi

Dokumen terkait

Pada hari ini Senin tanggal Empat bulan Juni tahun Dua ribu dua belas, selaku Panitia Pengadaan Barang/Pekerjaan Konstruksi dalam melakukan rapat Evaluasi

Unit Layanan Pengadaan Kota Banjarbaru mengundang penyedia Pengadaan Barang/Jasa untuk mengikuti Pelelangan Umum pada Pemerintah Kota Banjarbaru yang dibiayai dengan Dana APBD

dengan nenbawa surai iusas dli (di denFl L:sh) dari dircrtlr uha/Fimpitun peruehatr/kepara .ahrq dm kaiu pcns,enal5. s6eEry diradqr dMarni bbih daii r (etu) pqwahen

Model optimasi E conom ic Order Quantity dengan sistem parsial backorder dan increm ental discount dapat digunakan dalam masalah manajemen persediaan yang

Demikian pengumuman ini disampaikan untuk diketahui dan jika ada pihak-pihak yang berkeberatan dan atau merasa kurang puas terhadap hasil prakualifikasi ini dapat

Demikian undangan dari kami dan atas perhatiannya disampaikan terima kasih.. POKJA 1 ULP Kabupaten

diselenggarakan oleh satuan pendidikan yang terakreditasi atau lembaga sertifikasi.. Kewajiban akreditasi institusi sejak

Catatan : Agar membawa dokumen perusahaan asli sesuai dalam isian kualifikasi serta menyerahkan rekaman/copy-nyaM. Demikian undangan dari kami dan atas perhatiannya